Multiphoton Absorption of Myoglobin–Nitric Oxide Complex

Feb 22, 2012 - ... and Biochemistry, University of Texas at Austin, Austin, TX, United States ... V68W and L29F, illustrating a fair diversity of spat...
0 downloads 0 Views 5MB Size
Article pubs.acs.org/JPCB

Multiphoton Absorption of Myoglobin−Nitric Oxide Complex: Relaxation by D-NEMD of a Stationary State Grazia Cottone,*,†,§ Gianluca Lattanzi,†,∥ Giovanni Ciccotti,†,⊥ and Ron Elber‡ †

School of Physics, University College Dublin, Dublin, Rep. of Ireland Institute of Computational Engineering and Sciences (ICES) and Department of Chemistry and Biochemistry, University of Texas at Austin, Austin, TX, United States § Department of Physics, University of Palermo, Palermo, Italy ∥ Department of Basic Medical Sciences, University of Bari “Aldo Moro”, TIRES and INFN, Bari, Italy ⊥ Department of Physics, University of Rome “La Sapienza”, Rome, Italy ‡

ABSTRACT: The photodissociation and geminate recombination of nitric oxide in myoglobin, under continuous illumination, is modeled computationally. The relaxation of the photon energy into the protein matrix is also considered in a single simulation scheme that mimics a complete experimental setup. The dynamic approach to non-equilibrium molecular dynamics is used, starting from a steady state, to compute its relaxation to equilibrium. Simulations are conducted for the native form of sperm whale myoglobin and for two other mutants, V68W and L29F, illustrating a fair diversity of spatial and temporal geminate recombination processes. Energy flow to the heme and immediate protein environment provide hints to allostery. In particular, a pathway of energy flow between the heme and the FG loop is illustrated. Although the simulations were conducted for myoglobin only, the thermal fluctuations of the FG corner are in agreement with the large structural shifts of FG during the allosteric transition of tetrameric hemoglobin.

1. INTRODUCTION Sperm whale myoglobin is a small oxygen transport protein. It consists of 153 amino acids and a prosthetic heme group that binds small ligands (see Figure 1a). The physiological importance of myoglobin as a storage device of oxygen in muscles, and the empirical observation that it is stable and easy to handle experimentally, motivated the investigations of its structure, dynamics, and function. Indeed, it was the focus of a large number of biophysical studies (for a book, see ref 1). Of interest to the present manuscript are kinetic experiments. A typical kinetic setup includes the application of a flash that breaks the bond between a small ligand (L) and the iron atom of the heme (H).2−9 After dissociation of the ligand, the iron atom has only five coordinating groups instead of six and is displaced out of the heme plane (see Figure 2a). The shift of the iron position changes the geometry of the heme and is detected spectroscopically. Eventually, the heme acquires a replacement and reforms a bond with a ligand. The time scale and yield of bond reformation are recorded and used to fit rate coefficients and equilibrium constants for the process. Austin et al.4 pioneered the use of low temperature flash measurements and extended the experiments over a huge range of time scales. They observed the phenomenon of geminate recombination. A geminate recombination is the reformation of the L−H bond between the same pair that was bonded before the flash. In other words, the dissociated ligand does not leave the protein after the flash. It remains inside the protein matrix until activated diffusion brings it back to the heme group to © 2012 American Chemical Society

which it rebinds. Alternatively, the dissociated ligand may diffuse to the solvent in which it is lost (in practice) to the original heme group. Geminate recombination is particularly intriguing, since the dissociated ligand hops between cavities inside the protein matrix and the delay in the experimentally observed recombination can be associated with time lost in the cavities before returning to the binding site. Schematically, we can describe the process using the following kinetic scheme: kd

ks

L−H ⇌ L···H → L + H kr

(1)

The first state on the left (L−H) represents the ligand chemically bonded to the heme. The second, middle state, represents the geminate pair. The ligand in state (L···H) is still in the same protein matrix and did not have enough time to diffuse to the solvent. Finally, the state on the right (L + H) is that of a ligand that escapes its original protein and is now in aqueous solution. The probability of an escaping ligand to return back to the same protein molecule is usually negligible. Rate constants of eq 1 and variants of this equation are fitted to experimental data. Most efforts were invested in modeling one component of the geminate process, the middle step of ligand diffusion in the protein matrix. Tilton et al.10 used a probe Received: December 16, 2011 Revised: February 13, 2012 Published: February 22, 2012 3397

dx.doi.org/10.1021/jp212148x | J. Phys. Chem. B 2012, 116, 3397−3410

The Journal of Physical Chemistry B

Article

Figure 1. (a) Myoglobin in ribbon representation, with the indication of the four xenon cavities (transparent orange spheres). Elements of secondary structure (helices E, F, G, and H and loops CD and FG) are also indicated. (b) The heme pocket with the relevant residues lining the binding site (above the heme plane) and the proximal histidine (below the heme plane); the isopropionate side chains are also indicated. From left to right: native, V68W mutant, and L29F mutant (structures as in 1MBC.pdb, 20H9.pdb, and 2G0R.pdb). Images were prepared with the VMD program.48

Due to the loss of the diffusing ligand to solvent or to heme recombination, ligand diffusion is not an equilibrium process. Furthermore, the properties of the diffusive ligand are probed in the above recombination experiments indirectly, by fitting a nonlinear model of the kinetics to rebinding curves. In addition to kinetic curves of flash experiments, information on the average position of the ligand as a function of time is now available from time-resolved crystallography.18−20 These beautiful experiments provide “real life” images of ligand diffusion pathways. However, they do not provide new information on kinetics and thermodynamics. The experiments confirmed diffusion pathways and the existence of empty cavities detected earlier by the simulations of Tilton et al.10 and Elber and Karplus.11 We can therefore conclude that the locations of plausible diffusion pathways and protein cavities, that delay ligand rebinding and/or escape, are explored adequately. The kinetic parameters of the process are still under debate. Different computations provide results with mild variations.6 Another important component of the process described by eq 1 is that of energy relaxation. The dissociation process pumps energy to the ligand and to the heme. The excess energy

Figure 2. Sketch of the five- (a) and six-coordinated (b) heme in Mb. The four pyrrole N atoms, the Fe atom, and the closest nitrogen atom of the proximal hystidine (called Nε) are indicated.

ligand to explore empty spaces in myoglobin and suggested plausible diffusion pathways. Elber and Karplus11 used a mean field theory to explore diffusion pathways of a diatomic ligand in a thermally fluctuating myoglobin. More recent investigations either applied molecular dynamics simulations directly12,13 or estimated free energy surfaces for ligand diffusion with modern techniques such as TAMD,14 metadynamics,15,16 or particle insertion.17 3398

dx.doi.org/10.1021/jp212148x | J. Phys. Chem. B 2012, 116, 3397−3410

The Journal of Physical Chemistry B

Article

The nonequilibrium investigation described above is also useful for the investigation of the allosteric effect in hemoglobin. Hemoglobin is a tetramer, and each of the four globin chains (2 α chains and 2 β chains) resembles myoglobin. Early events of the allosteric process include dissociation of the ligand−heme bond and transmission of excess energy from the heme to the globin surface. Hence, the first step is internal and is unlikely to require communication between chains, making the myoglobin system a plausible model. By generating a steady state of photodissociated molecules, a significant amount of energy is pumped into the heme and its immediate environment. The excess energy is likely to provide a stronger signal to the mechanism of allosteric initiation and energy flow to the globin surface. The paper is organized as follows. In section 2, we describe the model for the geminate recombination under continuous illumination; in section 3, we briefly describe the dynamic nonequilibrium molecular dynamics framework. In section 4, the computational details are fully described. Results from the application to wild type myoglobin and two selected mutants are presented in section 5. Finally, further steps are discussed in the Conclusions.

dissipates, and the flow of energy is of particular interest, since it may be relevant to allosteric transitions in general and to the allosteric transition in hemoglobin in particular. While myoglobin is not an allosteric protein, the related protein hemoglobin is. Hemoglobin has four subunits, each subunit being highly homologous to myoglobin. The dissociation of a ligand in hemoglobin (by a flash or thermal activation) promotes an allosteric effect in which the relative packing of the subunits changes dramatically.21 The mechanism by which the dissociation of the ligand promotes the large structural change (the allosteric effect) is not known quantitatively. Myoglobin suggests a simple model system for the transfer of the signal (and energy) from the so-called allosteric core,22 of a ligand bound to a heme, to the interface of the interacting subunits. The study of the spatial and temporal energy relaxation after bond dissociation is therefore of considerable interest. Advanced spectroscopic techniques23−28 measure the rate at which the heme group loses its excess energy to the solvent. A computational model that explores geminate recombination of the ligand should also account correctly for the measured time scale of energy flow. Other investigators studied computationally energy relaxation and flow from excited heme.29−32 In particular, Sagnella and Straub32 carefully evaluated many aspects of this process. They proposed an intriguing mechanism in which excess energy is passed to the isopropionate side chains of the heme (see Figure 1b) and from there transported to the solvent. This mechanism was supported by an experimental study,27 in which the isopropionate side chains were removed and the difference in the relaxation was recorded. Finally, the dynamics of bond formation and/or dissociation of a small ligand and the heme iron has been investigated in previous works.8,33−35 It is therefore a valid question, why do we want to return to model myoglobin? The motivation behind the present study is that these components were not put together in a single comprehensive simulation scheme, in which multiple measurements are brought together and accounted for in a single framework. Moreover, in the present paper, we probe geminate recombination in a stationary nonequilibrium state. We consider a ligand (nitric oxide or NO) that can rebind to the heme iron extremely quickly (on the picosecond time scale). This time is comparable to the time scale of heme relaxation. The two processes are therefore coupled in time and presumably also in space. This is in contrast (for example) with the photodissociation of carbon monoxide for which the recombination time scale can exceed 100 ns. Another motivation to study NO is that nitric oxide binding is relevant to signaling processes in biology. Experiments can be done with continuous illumination of the sample (or using long nanosecond pulses26). In that case, if the light is intense enough, nitric oxide may have sufficient time to absorb a photon, dissociate, rebind, and repeat the cycle numerous times. The continuous illumination creates a stationary nonequilibrium state of dissociated and bound heme−iron nitric oxide pairs. The rapid rebinding of NO and its wide interest makes it well suited for non-equilibrium molecular dynamics simulations. We therefore concentrate in this paper on the geminate recombination of nitric oxide to myoglobin. Such a stationary system should be studied with the tools of nonequilibrium molecular dynamics.36−40 Once the photon beam is turned off, the system relaxes to its equilibrium state, starting however from the stationary nonequilibrium initial conditions.

2. THE MODEL In the present manuscript, we describe photodissociation and recombination processes of a small molecule (nitric oxide) with a heme group embedded in the protein myoglobin. To describe these molecular events, we need at least two Born− Oppenheimer energy surfaces: the ground state and an excited state (reachable with the exciting photon), and a mechanism for transitions between the two surfaces. The motions on each of these surfaces are handled by standard molecular dynamics. We describe below the energy surfaces first, and then the switching mechanisms. The calculations are conducted with the molecular modeling program MOIL41 in which the algorithms discussed were implemented in full. MOIL is available for download from http://clsb.ices.utexas.edu/prebuilt/. The current model follows ref 33 with only one significant addition. Previously, we allowed for only a single event of photon absorption. Here, we consider multiple occurrences of photon absorption, dissociation, and geminate recombination in a single trajectory. While the changes may seem small, they are significant. They allow us to explore the buildup of energy at the heme, examine multiple and correlated absorption of photons, and open the way for better description of nonequilibrium stationary states in biology and their relaxation. For completeness, we briefly describe the model. We refer the reader to the earlier publication33 for a list of the potential parameters. 2.1. Energy Surfaces. In molecular modeling programs, the interactions between the atoms (protein, ligand, and solvent atoms) are typically described by a single Born−Oppenheimer (BO) energy surface. The single energy surface is usually represented by a sum of mechanical energy terms U=

∑ Eb + ∑ Eθ + ∑ Eϕ + ∑ Eiϕ + ∑ E vdw θ

b

+

ϕ



vdw

∑ Eelec elec

(2)

where the terms (from left to right) are a summation of bonds, bond angles, torsions, improper torsions, van der Waals, and electrostatic interactions. 3399

dx.doi.org/10.1021/jp212148x | J. Phys. Chem. B 2012, 116, 3397−3410

The Journal of Physical Chemistry B

Article

In the present paper, we define two diabatic surfaces. The two surfaces are assumed to cross at a given distance between the iron atom and the nitrogen atom of the ligand. Most of the contributions to the energy surfaces follow eq 2. Moreover, a significant part of the contributions to the energy surfaces (e.g., those coming from water and the intraprotein parts which are outside the reaction site) is identical. The first diabatic surface, which we denote by Ud1, is the Born−Oppenheimer surface which corresponds to the ground state when the ligand, close to the iron, is bound to the heme. It is the state in which the system absorbs the photon. The second diabatic surface Ud2 is the surface that is reached once the photon is absorbed. We write Ud1 = Ud1,P − L + Ud1,P − H + Ud1,Fe − L

(3)

Ud2 = Ud2,P − L + Ud2,P − H + Ud2,Fe − L

(4)

The remaining diabatic terms represent the interaction of the ligand with the heme iron. The term Ud1,Fe−L is given by eq

Ud1,Fe − L = D{exp[−2α(RFe,N − RFe,N)] eq

− exp[−α(RFe,N − RFe,N)]}

The energy for heme−nitric oxide binding D is 30 kcal/mol; the Morse potential of eq 8 allows for asymptotic bond eq breaking at a cost of 30 kcal/mol. The parameters α and RFe,N eq were described in ref 33 and are α = 2 Å−1 and RFe,N = 1.743 Å. As for the term Ud2,Fe−L, it is given by Ud2,Fe − L = A exp( −bRFe,N) − B

P = exp( −πΓLZ/2)

ΓLZ =

(5)

where the parameter values are Rcut = 3 Å and λ = 0.2 Å, so that Ud1,P − H = [1 − f (RFe,N)]U6,P − H + f (RFe,N)U5,P − H (6)

The switching function changes from one at large iron− nitrogen distances to exp(−λ−1Rcut) = exp(−15) = 3.05 × 10−7 ≈ 0 at zero distance. In our model, the second diabatic surface Ud2,P−H is Ud2,P − H = U5,P − H

2πΔ2 hv|F1 − F2|

(10)

(11)

where Δ is the electronic coupling between the energy surfaces, h the Planck constant, v the modulus of the velocity of the reaction coordinate (the iron−nitrogen distance), and F1 and F2 the spatial derivatives of the two diabatic energy surfaces at the crossing point. The formula makes intuitive sense as follows. If the electronic coupling is zero, no transition is expected between the diabatic surfaces. Similarly, if the velocity is high, the transition probability between diabatic surfaces will be smaller, since the time the system spends at the transition domain is shorter. The second mechanism allows a bidirectional process which gets and releases energy. It is not uncommon to find the system moving on the left of the crossing point along the dissociated curve, crossing and changing temporarily to the associated diabatic surface, so gaining energy, then pass the crossing point

1 1 + exp[−λ−1(R − R cut)]

(9)

where A, B, and b are constants and RN−Fe is the distance between the iron and the nitrogen atom. 2.2. Switching between Electronic Curves. There are two mechanisms for the system to switch between the two diabatic energy surfaces. The first is by a direct absorption of a photon, and the second is by curve crossing. In the first mechanism, an absorbed photon transition takes the system from the ground to the excited state without a change in the nuclear positions or velocities. Within our model, it means that the energy function [Ud1,P−H + Ud1,Fe−L] is instantaneously changed to [Ud2,P−H + Ud2,Fe−L]. The absorption of the photon is restricted by the following conditions: (i) The ligand position should be such that its energy given by eq 8 is at most −20 kcal/mol; otherwise, we anticipate poor Franck−Condon overlap for the transition. (ii) There is a constant probability, p, for the molecule to absorb a photon in a time step. This probability depends (of course) on the concentration of myoglobin in the sample and on the intensity of the photons in the incoming beam. For computational convenience, we set this probability to be high and we experimented with photon-absorption probabilities of 0.1 and 0.05 photons/ps without significant changes in the qualitative behavior of the simulation. The second mechanism models a transition between diabatic curves with the Landau−Zener mechanism.44,45 We consider the probability P of the system to remain on the diabatic curve while passing through a crossing point. The crossing in our model is of the repulsive curve of eq 9 and the Morse potential of eq 8. The probability is given by

The notation P is for the protein, H for the heme, L for the ligand, and Fe for the iron. For simplicity, we are omitting the contributions coming from the solvent terms. Ud1,P−L and Ud2,P−L coincide and represent all the interaction terms in myoglobin (except the heme), and between myoglobin and the ligand atoms. Their functional form is standard. The energy parameters are from the OPLS force field with the united atom model42 (CHn groups are modeled as point masses) as implemented in the MOIL program.41 The NO ligand was modeled with the two-site simple model. It consists of LennardJones sites and small charges (0.028 au) of opposite sign, located at the ligand atomic positions. The water model is TIP3P.43 The second and third terms describe respectively the interactions of the protein with the heme and the iron with the ligand, in the two diabatic states. Ud1,P−H depends on the distance between the iron and the ligand. Therefore, it has to represent the heme coordinated to five or six nitrogen atoms, U5,P−H and U6,P−H, respectively. In contrast, Ud2,P−H represents always states with 5-coordinated heme, and therefore is equal to U5,P−H for all iron−ligand distances. In U5,P−H, the iron is slightly displaced from the heme plane in the direction of the proximal hystidine (see Figure 2a). The iron in U6,P−H is coordinated to six nitrogen atoms, including the one of the ligand; the equilibrium length of the bond between the Nε atom of the proximal histidine and the iron is changing from 2.1 Å in the 5-coordination state to 2.2 Å in the 6-coordination state (see Figure 2b). For a complete set of parameters, see ref 33. The term Ud1,P−H has a smoothed piecewise definition from 6-coordinated to 5-coordinated heme, given by a switching function f(R): f (R ) =

(8)

(7) 3400

dx.doi.org/10.1021/jp212148x | J. Phys. Chem. B 2012, 116, 3397−3410

The Journal of Physical Chemistry B

Article

microscopic observable Ô (Γ(t)). The time-dependent behavior of the macroscopic observable O(t) can be estimated by taking the arithmetic average of Ô (Γ(t)) over the trajectories originated from each of the chosen initial states which represent a sample from w(Γ, 0).

again and switch back to the excited state. Multiple cycles of dissociation and rebinding lead to further energy flow into and out of the heme. Multiple peaks in the heme kinetic energy as a function of time (see, for instance, Figures 6−8) are not necessarily following events of photon absorption. This complex nonmonotonic behavior of energy flow and significantly longer relaxation times were not observed in single photon-absorption events. They lead, as we illustrate in section 5, to the observation of new phenomena.

4. COMPUTATIONAL METHODS Molecular dynamics simulations were performed on three systems: the native form of myoglobin and two myoglobin mutants L29F and V68W in aqueous solution (PDB codes 1MBC,46 2G0R,18 and 2OH9,47 respectively). Figure 1a shows a ribbon representation of the protein and the center of the four xenon cavities. Figure 1b shows the heme pocket with the key residues lining the binding site on the distal side, in the native protein and in the L29F and V68W mutants. The figures were prepared by VMD.48 It is well-known that the L29F mutation, in which a leucine residue is replaced by the bulky group of phenylalanine (Figure 1b, right panel), shows biphasic kinetic behavior: at short times (picoseconds), the ligand is unable to escape from the iron atom and it recombines rapidly; ligands that are able to leave the heme pocket face a significantly longer time scale of recombination (nanoseconds, for CO); see ref 47 and references therein. Time-resolved X-ray experiments49 illustrated that the ligand stays in the binding site (above the heme) immediately after photolysis. After a while (